You are on page 1of 9

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/328190184

Effect of a soccer ball’s surface texture on its aerodynamics and trajectory

Article  in  Proceedings of the Institution of Mechanical Engineers Part P Journal of Sports Engineering and Technology · October 2018
DOI: 10.1177/1754337118794561

CITATIONS READS

5 303

3 authors:

Sungchan Hong John Eric Goff


University of Tsukuba University of Lynchburg
75 PUBLICATIONS   258 CITATIONS    42 PUBLICATIONS   385 CITATIONS   

SEE PROFILE SEE PROFILE

Takeshi Asai
University of Tsukuba
215 PUBLICATIONS   1,370 CITATIONS   

SEE PROFILE

Some of the authors of this publication are also working on these related projects:

Sports Ball Aerodynamics View project

Teaching Sports Physics / Sports Engineering View project

All content following this page was uploaded by Sungchan Hong on 19 June 2020.

The user has requested enhancement of the downloaded file.


Original Article

Proc IMechE Part P:


J Sports Engineering and Technology
1–8
Effect of a soccer ball’s surface texture Ó IMechE 2018
Article reuse guidelines:
on its aerodynamics and trajectory sagepub.com/journals-permissions
DOI: 10.1177/1754337118794561
journals.sagepub.com/home/pip

Sungchan Hong1 , John Eric Goff2 and Takeshi Asai1

Abstract
The effect of a soccer ball’s surface texture on its aerodynamics and flight trajectory is not definitively known. For this
study, five soccer balls were used, each having 32 panels with different surface textures. Their aerodynamics were exam-
ined via wind-tunnel experiments and then several non-spin trajectories were calculated for each ball. The results
showed that the aerodynamic forces acting on a soccer ball change significantly depending on the surface texture of the
ball, which in turn influences flight trajectories. The study contributes to an understanding of how a soccer ball’s surface
influences the aerodynamics, which may impact the future design and development of soccer balls.

Keywords
Aerodynamics, football, soccer, surface texture, trajectory

Date received: 26 September 2017; accepted: 24 July 2018

Introduction the EURO 2016 Cup. With only modest changes in


color, the World Cup ball was typically used as the
A traditional soccer ball is composed of 32 panels (12 official ball in league games that commenced after the
regular pentagonal panels and 20 regular hexagonal World Cup, as well as in the EURO Cup and the
panels). The ball’s surface structure has changed signifi- FIFA Confederations Cup. Since 2006, however, there
cantly over time with the number of panels being has been a slight change in the surface texture of the
reduced from 14, to 8, to 6. The ball’s trajectory has official ball (EURO PASS, 14-panel, adidas) used in
been shown1–3 to change with the number and orienta- the EURO Cup, which is held 2 years after the World
tion of the panels. Since the introduction of the official Cup. The official ball used through 2006 had a smooth
ball (Telstar, 32-panel, adidas) in the 1970 FIFA World surface. A soccer ball of modern design, having various
Cup held in Mexico, similar 32-panel balls have been protrusion patterns on the ball surface, was used in the
used as official balls of the World Cup, up to the EURO Cup of 2008. All the official balls of the EURO
Korea–Japan World Cup of 2002 (Fevernova, 32- Cup and World Cup since then have contained various
panel, adidas). After the 2002 World Cup, the shape kinds of protrusions (see Figure 1 for photographic
and design of the ball surface underwent substantial images of sample surfaces for six balls).
changes with the present pattern quite different than a Previous studies on soccer balls used various meth-
traditional soccer ball. A 14-panel ball, the Teamgeist ods, such as wind-tunnel experiments and computa-
by adidas, was used as the official ball for the German tionally determined flight trajectories.4–11 Studies have
World Cup held in 2006. The eight-panel Jabulani (adi-
also looked at roughness and the influence of panel fea-
das) ball was used in the 2010 South Africa World
tures, such as number, shape, and orientation, on ball
Cup. The aerodynamics of the Jabulani ball was notice-
aerodynamics.12–17 There is, however, little
ably different from those of the conventional ball.1,2
Jabulani was replaced by the Tango 12, 32-panel, adi-
das ball for the EURO 2012 Cup. 1
Institute of Health and Sports Science, University of Tsukuba, Tsukuba,
In the 2014 Brazil World Cup, a ball consisting of Japan
2
six panels (Brazuca, adidas) was used as the official ball Department of Physics, Lynchburg College, Lynchburg, VA, USA
and was reported to have a more stable trajectory than
Corresponding author:
other official balls.1,2 As a result, a ball with similar Sungchan Hong, Institute of Health and Sports Science, University of
panels, but a slightly different surface texture (Beau Tsukuba, Tsukuba 305-8574, Japan.
Jeu, six-panel, adidas), was used as the official ball in Email: hong.sungchan.fu@u.tsukuba.ac.jp
2 Proc IMechE Part P: J Sports Engineering and Technology 00(0)

Figure 1. Changes in the surface texture of modern soccer balls ((a) 2006 Germany World Cup ball; smooth texture, (b) 2008
EURO Cup ball; small pimple texture, (c) 2010 South Africa World Cup ball; wavelike rectangle texture, (d) 2012 EURO Cup ball;
small pimple texture, (e) Brazil World Cup 2014 ball; wavelike small pimple texture, and (f) 2016 EURO Cup ball; small square
texture).

understanding of the effect of the ball’s surface texture


on its aerodynamics and flight trajectory.
This study focused on the influence of surface tex-
ture on the aerodynamics and flight trajectory of five
32-panel soccer balls with different surface textures. All
five balls were made of the same leatherette (synthetic
materials) with panels attached by thermal bonding.
The aerodynamic forces for each ball were measured
with a wind tunnel in two different orientations. No-
spin flight trajectories were calculated using wind-
tunnel data.

Figure 2. Wind-tunnel setup.


Wind-tunnel experiment
The closed (return-flow) type, low-speed, low-
turbulence wind tunnel located at the University of nozzle cross section to adjust the distance between the
Tsukuba (San Technologies Co., Ltd.; Figure 2) was nozzle and the ball to zero, the flow generated from the
used in this experiment. support sting and column may not have a direct effect
The maximum wind speed was 55 m/s with a wind on the flow around the ball. Furthermore, the length of
speed distribution within 60.5%. The degree of turbu- the sting used in this study was 0.8 m and its width was
lence was less than 0.1%. The nozzle size was 1.5 m 0.02 m. Moreover, because the effect of sting vibrations
3 1.5 m and the blockage of the measured soccer ball is reduced to a small value by placing the six-
was within 5% of the nozzle size. For example, when component force detector behind the soccer ball, the
the wind speed was set to 25 m/s, the measured mean magnitude of the sting forces was ignored in this study.
wind speed was 25.28 m/s with a standard deviation Wind speed ranged from 7 m/s (Re ’ 105) to 35 m/s
from the measurement position in the range of 20.46 to (Re ’ 5 3 105). Measurements were made for each of
0.45 and a wind speed distribution within 60.5%. the five types of balls shown in Figure 3 at wind-speed
Similarly, the degree of turbulence downstream of the intervals of 1 m/s for a duration of 10 s (1000 fps;
nozzle when the wind speed was 25 m/s was 0.05–0.06 10,000 samples). Each of the five balls was then rotated
or within approximately 60.1%. Thus, the error due to 180° about an axis perpendicular to the images in
the ball position was believed to have little effect on the Figure 3.
wind speed. Furthermore, in the measurement system The balls used in this study were classified into five
of this study, the dynamic pressure can be automatically types, from A to E, according to the surface texture of
measured at 0.1 Pa intervals using a Pitot-static tube the ball. The surface of ball A had small square protru-
placed above the measurement portion of the soccer sions in a wavelike formation, whereas ball B was com-
ball. In addition, because the position of the ball during posed of a dimpled pattern. The surface of ball C
the measurement procedure was set at the center of the consisted of small triangular protrusions, while ball D
Hong et al. 3

Figure 3. Surface textures of the soccer balls used in this study while in the 0° orientation. To obtain the 180° orientation, rotate
the images above by 180° about an axis perpendicular to the images (or flip this published page upside down). Ball A had waves of
square protrusions, ball B had a dimpled pattern, ball C had triangular protrusions, ball D had hexagonal textures, and ball E had
uniform squares.

had hexagons configured in the form of a honeycomb. detector (LMC-61256, Nissho Electric Works).
The surface of ball E was made up of evenly arranged Aerodynamic forces drag (D), lift (L), and side (S) were
small rectangular protrusions. All balls had diameter converted to the drag coefficient (CD), lift coefficient
0.22 m and masses in the range 0.428 kg to 0.433 kg. (CL), and side force coefficient (CS), as given by equa-
The inner gauge pressure of all balls was 9.0 psi. Non- tions (1)–(3)
spinning balls were studied in the wind tunnel. As
shown in Figure 2, each ball had to be destroyed for 2D
CD ¼ ð1Þ
play when tested in a wind tunnel. All five balls had rv2 A
one of their regular hexagons facing the oncoming air
(see orientation in Figure 3) to ensure that the balls 2L
CL ¼ ð2Þ
were oriented in the same way in the wind tunnel. rv2 A
Because all the balls have a traditional 32-panel surface
structure, the surface texture was the focus of this
and
study. Differences in aerodynamic coefficients will
result from different features on the ball surfaces, not 2S
from the arrangement of panels. The authors are aware CS ¼ ð3Þ
rv2 A
that the choice of ball orientation in the wind tunnel is
one of the infinite number of possible ball orientations.
The balls used in this study were manufactured in-house respectively, where r = 1.2 kg/m3 is air density, v is the
and, as mentioned, had to be destroyed for play when air speed, and A is the cross-sectional area of the ball
placed in the wind tunnel. Changing ball orientation given by A = p 3 (0.11 m)2 = 0.038 m2.
requires the manufacture of five new balls, a new round
of wind-tunnel testing, and then trajectory analyses.
Although it would be ideal to have a large supply of
Wind-tunnel experimental results
balls to investigate many ball orientations, resources Figure 4 shows the drag coefficient for the five soccer
and time limited the extent of the research. We urge the balls analyzed in this study. The drag coefficient for
reader to keep in mind that the results presented in the balls B and E decreased more rapidly than for the other
next section are for two orientations of non-spinning balls. This could be because the dimpled texture of ball
balls. This study represents a first pass at addressing the B’s surface and the uniform squares of ball E move the
contribution of surface texture on a soccer ball’s flight. separation of the boundary layer to the back of the ball
The aerodynamic forces acting on the soccer ball faster than for the other balls, hastening the transition
were measured using a sting-type, six-component force from a laminar boundary layer to a turbulent one. For
4 Proc IMechE Part P: J Sports Engineering and Technology 00(0)

Figure 4. Drag coefficient versus air speed for five balls under investigation. Symbols represent actual data; lines between symbols
are merely to guide the eye. The 0° orientation is on the left and the 180° orientation is on the right.

Table 1. Critical speeds, critical Reynolds numbers, and corresponding drag coefficients at the critical speed for the five balls at the
two tested orientations.

Ball vc (m/s) at 0° Rec 3 1025 at 0° CD(vc) at 0° vc (m/s) at 180° Rec 3 1025 at 180° CD(vc) at 180°

A 18.8 2.63 0.113 17.9 2.50 0.122


B 18.8 2.58 0.128 17.9 2.44 0.136
C 17.8 2.39 0.124 17.8 2.38 0.119
D 18.8 2.53 0.112 18.8 2.53 0.131
E 16.8 2.29 0.143 17.9 2.43 0.157

ball speeds mainly associated with passes and long Differences in Reynolds number at the same wind
kicks, the drag curve is in the critical and supercritical speed are due to temperature differences. Experiments
regions with speeds between approximately 10 m/s were performed in the temperature range of 28 °C–
(Re’ 1.4 3 105) and 30 m/s (Re’ 4.3 3 105). Note that 35 °C. Reynolds number is proportional to kinematic
the Reynolds number is defined as Re = vD/n, where v viscosity, which increases with increasing tempera-
is ball speed, D = 0.22 m is ball diameter, and n = 1.54 ture. Because drag coefficients could not be exactly
3 1025 m2/s is the kinematic viscosity at 24 °C. reproduced when balls were rotated 180°, the best
Note that a 180° rotation of the balls in Figure 3 is conclusions that may be drawn from Table 1 are that
not a symmetric rotation. Such a rotation effectively ball E with its uniform squares had the largest drag
interchanges the pentagons and hexagons that border coefficient at the critical speed and ball B with its
the forward-facing hexagon. Such a rotation should dimpled pattern had the second largest drag coeffi-
not, in principle, alter the drag force on ball. Figure 4 cient at the critical speed. Figure 4 shows that ball D
shows slightly different drag coefficients for the two with its hexagonal textures had the largest high-speed
orientations. Rotating a ball by 180° is more compli- (v . 32 m/s) drag coefficient; ball C with its triangu-
cated than it may seem because the fitting mechanism lar protrusions had the smallest high-speed drag coef-
that ensures the ball is attached to the support rod must ficient. Ball A with its waves of square protrusions
be rotated as well. It is thus quite difficult to ensure a had a drag coefficient curve that changed the most
perfect replication of the first experiment when a ball is when rotated by 180°.
rotated for the second experiment. Figure 5 shows wind-tunnel results for the lift coeffi-
The critical speed, vc, is the speed at which the drag cient for the five balls with different surfaces at each of
coefficient is at its minimum value.15 As air speed is the two tested orientations. The interchange of penta-
reduced, the drag coefficient rises precipitously in the gons and hexagons when balls are rotated 180° leads to
critical region. The transition from supercritical to sub- lift coefficients that do not simply switch signs, as
critical represents a change in air flow about the ball would be the case if each panel was the same geometric
from turbulent to laminar. Critical speeds, correspond- shape. All balls had roughly the same lift coefficient at
ing critical Reynolds numbers (Rec), and values of the the top speed in the wind tunnel, but ball A’s lift coeffi-
drag coefficient at the critical speed are all given in cient was the largest magnitude for most speeds above
Table 1. 20 m/s.
Hong et al. 5

Figure 5. Lift coefficient versus air speed for five balls. Symbols represent actual data; lines between symbols are merely to guide
the eye. The 0° orientation is on the left and the 180° orientation is on the right.

Figure 6. Side coefficient versus air speed for five balls. Symbols represent actual data; lines between symbols are merely to guide
the eye. The 0° orientation is on the left and the 180° orientation is on the right.

Figure 6 shows the speed-dependent side coefficient the left and right points of the center-facing hexagons
of the five balls under consideration for each orienta- in the photos shown in Figure 3. A 90° rotation of the
tion. Like the lift coefficient, a 180° ball rotation should balls in those photos leads to the flat sides of the hexa-
simply switch the sign of the side coefficient. Also like gons being on the left and right sides, which means
the lift coefficient, such a switch is not exactly seen in there are no longer seams on the left and right sides.
Figure 6, but the trend is visible. Balls A, B, and E Hexagons have rotational symmetry if the rotation is
showed relatively large side coefficient magnitudes for 60°, not 90°.
speeds larger than the critical speed. At the largest One may wonder why lift and side coefficients are
wind-tunnel speed, ball D had the largest magnitude of not zero for all speeds. The orientation of the hexagons
side coefficient. There was significant variation in side in Figure 3 suggests that if the balls are perfectly made
force applied to the balls according to the ball type. and perfectly aligned in the wind tunnel, there should
Because the lift and side coefficients originate from the be no sideways asymmetric shedding of the boundary
same physical phenomenon, namely, an asymmetric layer. But perfect alignment in the wind tunnel and per-
separation of the boundary layer from the back of the fectly symmetric surface textures are both impossible to
ball, rotation of a ball in the wind tunnel by 90° will attain. These results show that even slight variations
interchange lift and side coefficients (up to a sign, which from symmetric wind-tunnel placement and symmetric
depends on the direction of rotation). All such a rota- surface textures lead to nontrivial transverse aerody-
tion would do is interchange what is meant by sideways namic forces. Ball manufacturers should take heed of
and upward. The reason Figures 5 and 6 do not look the aforementioned result. No matter how well a ball is
the same is because a 90° rotation of the balls shown in made and how much effort is put into making a ball as
Figure 3 does not lead to the same configuration shown symmetric as possible for certain orientations, the like-
in Figure 3. Note the presence of seams coming from lihood of knuckle effect is high.
6 Proc IMechE Part P: J Sports Engineering and Technology 00(0)

Figure 7. Force variation in lift and side forces due to the wind speed.

Figure 8. Numerically calculated horizontal ranges for the five tested balls. Symbols are merely meant to guide the eye because
trajectories may be determined for any speed in the range of launch speeds (15–35 m/s). Balls were launched at 25° above the
horizontal in the computer code. The 0° orientation is on the left and the 180° orientation is on the right.

Figure 7 shows the graph of the force variation condition within a speed range (15–35 m/s) that repre-
(standard deviation (SD)) due to the increase in the sents the majority of soccer kicks for which aerody-
flow velocity on each ball for 10 s. First, the lift force namic forces play significant roles in the trajectories.
and side force acting on the ball are seen to vary greatly The aerodynamic coefficients are taken to be as they
depending on the type of ball. Furthermore, although appear in Figures 4–6, which means linear interpolation
the force variation is different for different types of is used between wind-tunnel data. Non-spin trajectories
balls, for ball B at flow velocity 30 m/s, the variation were numerically determined for balls launched at 25°
exceeded 2.5 N in terms of both the lift and side forces, above the horizontal with speeds ranging from 15 to
showing a greater variation than that for the other 35 m/s. Data were collected for two ball orientations
types of balls with an increase in flow velocity. and the ball orientation was assumed to remain fixed
However, for ball D, the force variation was less than during the flights in the model trajectories. By analyzing
1.5 N, showing a smaller variation than that for the non-spin trajectories, the focus is on balls which exhibit
other balls. For example, in the case of the flow velo- knuckle effects.18 The authors’ previous work with non-
city at 30 m/s, the variation in the lift and side forces spinning ball trajectory19 provided insight regarding
was 2.6 and 2.6 N, respectively, for the B type ball and examination of trajectory results for this study.
1.0 and 1.5 N, respectively, for the D type ball. Figure 8 shows the horizontal range for the trajec-
tories. Because the surfaces of balls A and E lead to
larger magnitude lift coefficients at post-critical speeds
Trajectories
(see Figure 5), balls A and E showed the most influence
Armed with all aerodynamic coefficients as functions of from nonzero lift coefficients. The ranges for balls A
speed for the five balls, methods developed elsewhere13 and E were approximately 15 and 7.5 m different than
were employed to evaluate trajectories for any launch the ranges for balls B, C, and D.
Hong et al. 7

Figure 9. Numerically calculated lateral deflections as percentages of horizontal ranges for the five tested balls. Symbols are merely
meant to guide the eye because trajectories may be determined for any speed in the range of launch speeds (15–35 m/s). Balls were
launched at 25° above the horizontal in the computer code. The 0° orientation is on the left and the 180° orientation is on the right.

Figure 10. Numerically calculated effects on horizontal ranges for the five tested balls when the lift coefficient is set to zero.
Symbols are merely meant to guide the eye because trajectories may be determined for any speed in the range of launch speeds
(15–35 m/s). Balls were launched at 25° above the horizontal in our computer code. The 0° orientation is on the left and the 180°
orientation is on the right.

Figure 9 shows the fraction of the horizontal range authors’ conclusions on how the surfaces of balls A
represented by the lateral deflection for each of the and E influence their trajectories require further testing
numerically determined trajectories. Figure 6 explains in future work that will focus on flow visualization
why balls D and E showed the most lateral deflection across the different surfaces.
as percentages of the range. For most post-critical
speeds, ball E showed a lateral deflection of 13%–18%.
Such large lateral deflections make goal keeper’s jobs
Conclusion
much harder when trying to stop non-spinning balls In this study, the influence of the surface texture of the
from entering the goal. Figure 8 shows that if ball E is soccer ball on its aerodynamics and flight characteris-
launched at 30 m/s, it could travel at least 40 m. A 15% tics were examined by carrying out wind-tunnel experi-
lateral deflection amounts to 6 m. ments on five geometrically identical 32-panel balls
Figure 10 shows how the horizontal range is chan- having different surface textures. The results indicate
ged when the lift coefficient is set to zero in the numeri- that the pattern on the surface of a soccer ball signifi-
cally determined trajectories. As suggested by Figure 8, cantly influences the ball’s trajectory. Balls A and E
ranges for balls A and E showed the greatest depen- with their pimpled protrusions and squares, respec-
dence on their lift coefficients. Ball A’s range may be tively, had surfaces that influenced numerically deter-
altered as much as 30% by its nonzero lift coefficient, mined trajectories more than dimpled surfaces on the
another effect caused by the protrusions on ball A. The other balls tested. Goal keepers will likely be challenged
8 Proc IMechE Part P: J Sports Engineering and Technology 00(0)

by ball trajectories with lateral deflections as large as 2. Hong S and Asai T. Effect of panel shape of soccer ball
10%–20% of a ball’s range. Although these wind- on its flight characteristics. Sci Rep 2014; 4: 5068.
tunnel tests only examined one portion of a ball facing 3. Goff JE, Hong S and Asai T. Aerodynamic and surface
the oncoming air, meaning the trajectory analyses comparisons between Telstar 18 and Brazuca. Proc
could only examine non-spinning balls of a single IMechE, Part P: J Sports Engineering and Technology.
orientation throughout flight, the authors have deter- Epub ahead of print 19 May 2018. DOI: 10.1177/
1754337118773214.
mined the effects of the lift and side forces on trajec-
4. Alam F, Chowdhury H, Moria H, et al. Aerodynamics
tories. Selection of soccer ball surface type will greatly of contemporary FIFA soccer balls. Procedia Engineer
influence trajectories of non-spinning balls in matches. 2014; 13: 188–193.
If event organizers are interested in minimizing erratic 5. Alam F, Chowdhury H, Stemmer M, et al. Effects of sur-
flights of non-spinning balls, the initial recommenda- face structure on soccer ball aerodynamics. Procedia
tion is to employ balls with dimpled surface features, Engineer 2012; 34: 146–151.
instead of pimpled surface features. ‘‘Erratic’’ flights 6. Asai T, Ito S, Seo K, et al. Characteristics of modern
could be interpreted as trajectories that are unpredict- soccer balls. Procedia Engineer 2012; 34: 122–127.
able from a player’s point of view, not trajectories that 7. Asai T and Seo K. Aerodynamic drag of modern soccer
are unpredictable from the physical forces on the ball. balls. SpringerPlus 2013; 2: 171.
Future work should assess the aerodynamics of spin- 8. Asai T, Seo K, Kobayashi O, et al. Fundamental aerody-
ning balls to determine if the larger effects of protru- namics of the soccer ball. Sports Eng 2007; 10: 101–110.
9. Mizota T, Kurogi K, Ohya Y, et al. The strange flight
sions are averaged out or if the spin enhances those
behaviour of slowly spinning soccer balls. Sci Rep 2013;
effects. That work will have to be compared with
3: 1871.
recently published research20 on balls with both slow 10. Oggiano L and Sætran L. Aerodynamics of modern soc-
and fast spins. cer balls. Procedia Engineer 2010; 2: 2473–2479.
11. Mehta RD. Aerodynamics of sports balls. Annu Rev Fluid
Acknowledgements Mech 1985; 17: 151–189.
12. Passmore M, Rogers D, Tuplin S, et al. The aerodynamic
The authors thank the two paper reviewers for their performance of a range of FIFA-approved footballs.
excellent work in significantly improving the quality of Proc IMechE, Part P: J Sports Engineering and Technol-
our paper. ogy 2012; 226: 61–70.
13. Goff JE and Carré MJ. Trajectory analysis of a soccer
ball. Am J Phys 2009; 77: 1020–1027.
Declaration of conflicting interests
14. Hong S, Asai T and Seo K. Visualization of air flow
The author(s) declared no potential conflicts of interest around soccer ball using a particle image velocimetry. Sci
with respect to the research, authorship, and/or publi- Rep 2015; 5: 15108.
cation of this article. 15. Achenbach E. The effects of surface roughness and tunnel
blockage on the flow past spheres. J Fluid Mech 1974; 61:
113–125.
Funding 16. Carré MJ and Barber S. A novel sports ball aerodynamics
The author(s) received no financial support for the analysis tool: soccer ball design. Sports Technol 2012; 3:
research, authorship, and/or publication of this article. 240–252.
17. Barber S and Carré MJ. The effect of surface geometry
on soccer ball trajectories. Sports Eng 2010; 13: 47–55.
ORCID iD 18. Hong S, Chung C, Nakayama M, et al. Unsteady aerody-
Sungchan Hong https://orcid.org/0000-0003-0149- namic force on a knuckleball in soccer. Procedia Engineer
5028 2010; 2: 2455–2460.
19. Goff JE, Hobson CM, Asai T, et al. Wind-tunnel experi-
ments and trajectory analyses for five nonspinning soccer
References balls. Procedia Engineer 2016; 147: 32–37.
1. Goff JE, Asai T and Hong S. A comparison of Jabulani 20. Passmore MA, Tuplin S and Stawski A. The real-time
and Brazuca non-spin aerodynamics. Proc IMechE, measurement of football aerodynamic loads under spin-
Part P: J Sports Engineering and Technology 2014; 228: ning conditions. Proc IMechE, Part P: J Sports Engineer-
188–194. ing and Technology 2017; 231: 262–274.

View publication stats

You might also like